HIGH-PERFORMANCE SIGNAL ACQUISITION...

153
HIGH-PERFORMANCE SIGNAL ACQUISITION ALGORITHMS FOR WIRELESS COMMUNICATIONS RECEIVERS A Dissertation by KAI SHI Submitted to the Office of Graduate Studies of Texas A&M University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY August 2005 Major Subject: Electrical Engineering

Transcript of HIGH-PERFORMANCE SIGNAL ACQUISITION...

  • HIGH-PERFORMANCE SIGNAL ACQUISITION ALGORITHMS

    FOR WIRELESS COMMUNICATIONS RECEIVERS

    A Dissertation

    by

    KAI SHI

    Submitted to the Office of Graduate Studies ofTexas A&M University

    in partial fulfillment of the requirements for the degree of

    DOCTOR OF PHILOSOPHY

    August 2005

    Major Subject: Electrical Engineering

  • HIGH-PERFORMANCE SIGNAL ACQUISITION ALGORITHMS

    FOR WIRELESS COMMUNICATIONS RECEIVERS

    A Dissertation

    by

    KAI SHI

    Submitted to the Office of Graduate Studies ofTexas A&M University

    in partial fulfillment of the requirements for the degree of

    DOCTOR OF PHILOSOPHY

    Approved by:

    Chair of Committee, Erchin SerpedinCommittee Members, Scott L. Miller

    Xi ZhangAndreas Klappenecker

    Head of Department, Chanan Singh

    August 2005

    Major Subject: Electrical Engineering

  • iii

    ABSTRACT

    High-Performance Signal Acquisition Algorithms

    for Wireless Communications Receivers. (August 2005)

    Kai Shi, B.Sc., Nanjing University, Nanjing, China;

    M.Eng., Southeast University, Nanjing, China

    Chair of Advisory Committee: Dr. Erchin Serpedin

    Due to the uncertainties introduced by the propagation channel, and RF and

    mixed signal circuits imperfections, digital communication receivers require efficient

    and robust signal acquisition algorithms for timing and carrier recovery, and interfer-

    ence rejection.

    The main theme of this work is the development of efficient and robust signal

    synchronization and interference rejection schemes for narrowband, wideband and

    ultra wideband communications systems. A series of novel signal acquisition schemes

    together with their performance analysis and comparisons with existing state-of-the-

    art results are introduced. The design effort is first focused on narrowband systems,

    and then on wideband and ultra wideband systems.

    For single carrier modulated narrowband systems, it is found that conventional

    timing recovery schemes present low efficiency, e.g., certain feedback timing recov-

    ery schemes exhibit the so-called hang-up phenomenon, while another class of blind

    feedforward timing recovery schemes presents large self-noise. Based on a general re-

    search framework, we propose new anti-hangup algorithms and prefiltering techniques

    to speed up the feedback timing recovery and reduce the self-noise of feedforward tim-

    ing estimators, respectively.

    Orthogonal frequency division multiplexing (OFDM) technique is well suited for

    wideband wireless systems. However, OFDM receivers require high performance car-

  • iv

    rier and timing synchronization. A new coarse synchronization scheme is proposed for

    efficient carrier frequency offset and timing acquisition. Also, a novel highly accurate

    decision-directed algorithm is proposed to track and compensate the residual phase

    and timing errors after the coarse synchronization step. Both theoretical analysis

    and computer simulations indicate that the proposed algorithms greatly improve the

    performance of OFDM receivers.

    The results of an in-depth study show that a narrowband interference (NBI) could

    cause serious performance loss in multiband OFDM based ultra-wideband (UWB) sys-

    tems. A novel NBI mitigation scheme, based on a digital NBI detector and adaptive

    analog notch filter bank, is proposed to reduce the effects of NBI in UWB systems.

    Simulation results show that the proposed NBI mitigation scheme improves signifi-

    cantly the performance of a standard UWB receiver (this improvement manifests as

    a signal-to-noise ratio (SNR) gain of 9 dB).

  • v

    To my parents and wife

  • vi

    ACKNOWLEDGMENTS

    I would like to thank my advisor, Professor Erchin Serpedin, for his constant

    support, guidance and inspiration during my years at TAMU. This research work

    would never have been accomplished without his encouragement and help.

    Special thanks goes to Prof. Scott L. Miller for teaching me modulation theory

    and serving as a member of my Ph.D. committee. I would also like to thank Prof.

    Xi Zhang and Prof. Andreas Klappenecker for serving on my advisory committee. I

    am also thankful to some professors in our department, Prof. Krishna R. Narayanan

    for his lectures on coding theory, and Prof. Aydin Karsilayan for his inspiration and

    guidance when I was working on the UWB project.

    I also enjoyed working with many colleagues in the Wireless Communications

    Laboratory (WCL), and in particular I would like to thank Dr. Yang Wang, Yik-

    chung Wu, Dr. Zhongmin Liu, Dr. Shengjie Zhao, Dr. Jing Li, Dr. Guoshen Yue,

    Nitin A. Nangare, Janath Peiris, Dr. Vivek Gulati, Hari Sankar, Yong Sun and Jun

    Zheng. I owe also special thanks to some researchers from the Analog and Mixed

    Signal Group at TAMU: Burak Kelleci, Timothy Fischer, and Haitao Tong.

    I would like to thank Dr. Zoran Zvonar, Dr. Aiguo Yan, Deepak Mathew and

    Lidwine Martinot for their encouragement and guidance when I worked as a system

    engineering intern at Analog Devices. This successful experience will have a strong

    impact on my future career.

    Finally, I would like to express my gratitude to my parents and sister for their

    encouragement and constant support. This dissertation is dedicated to my wife,

    Likun Cao. Her love, understanding and encouragement represented the momentum

    to complete this Ph.D. program.

  • vii

    TABLE OF CONTENTS

    CHAPTER Page

    I INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . 1

    A. Background . . . . . . . . . . . . . . . . . . . . . . . . . . 1

    1. Timing Recovery . . . . . . . . . . . . . . . . . . . . . 3

    2. Carrier Recovery . . . . . . . . . . . . . . . . . . . . . 4

    3. Interference Mitigation . . . . . . . . . . . . . . . . . 5

    B. Organization of Dissertation . . . . . . . . . . . . . . . . . 5

    II FAST FEEDBACK TIMING RECOVERY FOR NARROW-

    BAND SYSTEMS . . . . . . . . . . . . . . . . . . . . . . . . . . 9

    A. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . 9

    B. Signal Model and Feedback Timing Recovery . . . . . . . . 10

    C. Hangup in Digital PLL . . . . . . . . . . . . . . . . . . . . 12

    D. Anti-hangup Timing Recovery Scheme . . . . . . . . . . . 14

    E. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

    III JITTER-FREE FEEDFORWARD TIMING RECOVERY FOR

    NARROWBAND SYSTEMS . . . . . . . . . . . . . . . . . . . . 25

    A. Jitter-Free Prefilter for Feedforward Timing Recovery

    of Linear Modulations . . . . . . . . . . . . . . . . . . . . 25

    1. Introduction . . . . . . . . . . . . . . . . . . . . . . . 25

    2. Signal Model and Symbol Timing Estimators . . . . . 26

    3. Nearly Jitter-Free Prefilter . . . . . . . . . . . . . . . 28

    4. Summary . . . . . . . . . . . . . . . . . . . . . . . . . 34

    B. Jitter-Free Prefilter for Feedforward Timing Recovery

    of GMSK Modulations . . . . . . . . . . . . . . . . . . . . 34

    1. Introduction . . . . . . . . . . . . . . . . . . . . . . . 34

    2. Signal Model and Estimators . . . . . . . . . . . . . . 35

    3. Self-Noise Compensating Prefilter . . . . . . . . . . . 39

    4. Summary . . . . . . . . . . . . . . . . . . . . . . . . . 42

    IV EFFICIENT COARSE FRAME AND CARRIER SYNCHRO-

    NIZATION FOR OFDM SYSTEMS . . . . . . . . . . . . . . . . 43

    A. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . 43

  • viii

    CHAPTER Page

    B. Signal Model . . . . . . . . . . . . . . . . . . . . . . . . . 45

    C. Maximum Likelihood Estimator . . . . . . . . . . . . . . . 47

    D. Theoretical Analysis of Estimators . . . . . . . . . . . . . 56

    1. Coarse Packet Detection . . . . . . . . . . . . . . . . . 56

    2. Fine Packet Detection . . . . . . . . . . . . . . . . . . 58

    3. Carrier Frequency Offset Estimation . . . . . . . . . . 60

    E. Numerical Analysis of Estimators . . . . . . . . . . . . . . 63

    1. Performance of Algorithms . . . . . . . . . . . . . . . 65

    2. System Performance . . . . . . . . . . . . . . . . . . . 69

    F. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . 70

    V DECISION-DIRECTED FINE SYNCHRONIZATION FOR

    OFDM SYSTEMS . . . . . . . . . . . . . . . . . . . . . . . . . . 71

    A. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . 71

    B. Signal Models . . . . . . . . . . . . . . . . . . . . . . . . . 71

    C. Data-aided Estimator . . . . . . . . . . . . . . . . . . . . . 73

    D. Proposed Synchronization Scheme . . . . . . . . . . . . . . 74

    1. Decision-Directed Estimator . . . . . . . . . . . . . . 74

    2. Closed-Loop Scheme . . . . . . . . . . . . . . . . . . . 77

    E. Computer Simulations . . . . . . . . . . . . . . . . . . . . 78

    F. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

    VI ADAPTIVE NARROWBAND INTERFERENCE MITIGA-

    TION FOR MB-OFDM UWB SYSTEMS . . . . . . . . . . . . . 80

    A. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . 80

    B. Signal Model . . . . . . . . . . . . . . . . . . . . . . . . . 83

    1. Single-Band OFDM Systems . . . . . . . . . . . . . . 83

    2. Multi-Band OFDM Systems . . . . . . . . . . . . . . 85

    C. Impact of Narrowband Interference . . . . . . . . . . . . . 86

    1. Optimization of Data Converters . . . . . . . . . . . . 86

    2. Carrier Synchronization . . . . . . . . . . . . . . . . . 92

    3. System Performance of a MB-OFDM Receiver . . . . 96

    D. Narrowband Interference Suppression . . . . . . . . . . . . 98

    1. Digital NBI Detection and Mitigation . . . . . . . . . 99

    2. Analog NBI Cancellation . . . . . . . . . . . . . . . . 102

    3. Performance of the Mixed NBI Suppression Scheme . 111

    E. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . 113

    VII CONCLUSIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . 115

  • ix

    CHAPTER Page

    REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116

    APPENDIX A . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126

    APPENDIX B . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130

    APPENDIX C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132

    APPENDIX D . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133

    APPENDIX E . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135

    APPENDIX F . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137

    VITA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139

  • x

    LIST OF TABLES

    TABLE Page

    I Optimum Ω versus ADC bit number for OFDM receiver in AWGN

    channels. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89

    II The simulation configuration of MB-OFDM systems. . . . . . . . . . 98

  • xi

    LIST OF FIGURES

    FIGURE Page

    1 Signal acquisition at digital receiver. . . . . . . . . . . . . . . . . . . 2

    2 Feedback timing recovery scheme. . . . . . . . . . . . . . . . . . . . . 3

    3 Feedforward timing recovery scheme. . . . . . . . . . . . . . . . . . . 4

    4 Coarse and fine synchronization of OFDM receiver. . . . . . . . . . . 5

    5 Content and organization of dissertation. . . . . . . . . . . . . . . . . 6

    6 Conventional timing recovery scheme. . . . . . . . . . . . . . . . . . 12

    7 The S-curves of timing error detectors. . . . . . . . . . . . . . . . . . 14

    8 Timing errors versus the number of symbols for Gardner’s TED. . . . 15

    9 The shifted version of S-curves. . . . . . . . . . . . . . . . . . . . . . 16

    10 The new S-curves. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

    11 Fast timing recovery scheme. . . . . . . . . . . . . . . . . . . . . . . 17

    12 Ph versus different L. . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

    13 Performance comparison in AWGN channels, Es/No = 15 dB, 100 runs. 21

    14 Performance comparison in flat fading channels, BLT = 0.01,

    Es/No = 20 dB, fdT = 0.0002. . . . . . . . . . . . . . . . . . . . . . . 23

    15 Plots of Gm(1, 0)/Gm(0, 0) for m = 1, 2 versus various roll-off

    factors ρ. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

    16 Comparison of symbol timing estimators for P = 2. . . . . . . . . . . 32

    17 Comparison of the performance of estimator for different N . . . . . . 33

    18 Feedforward clock recovery scheme. . . . . . . . . . . . . . . . . . . . 37

  • xii

    FIGURE Page

    19 Performance comparisons of various estimators for GMSK modulations. 38

    20 Amplitude response of filters. . . . . . . . . . . . . . . . . . . . . . . 40

    21 The performance of estimators with and without prefilter. . . . . . . 41

    22 Mean values of timing metrics, SNR=20dB. . . . . . . . . . . . . . . 52

    23 Coarse packet detection and fine packet detection(Ω = L). . . . . . . 55

    24 Probability of Z(l) > 0 for (a) Es/N0 = 10dB, (b) Es/N0 = 40dB.

    The solid and the dash lines are associated with the metrics of [+B

    +B -B +B] and [+A +A], respectively. . . . . . . . . . . . . . . . . . 57

    25 The probability of false alarm versus N and Tc. . . . . . . . . . . . . 59

    26 MSE of carrier frequency offset estimator in AWGN channels. . . . . 61

    27 Comparison of MSE and complexity of carrier frequency offset

    estimators in AWGN channels, SNR=15dB. . . . . . . . . . . . . . . 64

    28 Time offset estimation l̂ in frequency-selective channels. . . . . . . . 66

    29 The probabilities of timing errors that will cause ISI in frequency-

    selective channels. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

    30 MSE of the carrier frequency offset estimators in frequency-selective

    channels. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68

    31 BER performance versus SNR in frequency-selective channels. . . . . 69

    32 The receiver structure. . . . . . . . . . . . . . . . . . . . . . . . . . . 75

    33 Normalized MSE (normalized by 10−12) of open-loop DD SFO

    estimators. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

    34 Normalized MSE (normalized by 10−12) of closed-loop SFO estimators. 78

    35 The signal model of single-band OFDM systems. . . . . . . . . . . . 83

    36 γq versus Ω and number of bits of ADC. . . . . . . . . . . . . . . . . 88

  • xiii

    FIGURE Page

    37 SINR curves of ADC, Pc/Pn = 10dB. . . . . . . . . . . . . . . . . . . 91

    38 Theoretical (dashed line) and simulated (solid line) optimum γqversus SIRs. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92

    39 SINR with non-optimal γq (solid line) and with optimal γq (dashed

    line), Pc/Pn = 10dB. . . . . . . . . . . . . . . . . . . . . . . . . . . . 93

    40 The performance of CFO estimators versus SIR, v = 0.08, L =

    128, and Pc/Pn = 5dB. . . . . . . . . . . . . . . . . . . . . . . . . . . 96

    41 The block error rate versus Pc/Pn and SIR. . . . . . . . . . . . . . . 97

    42 The magnitude-squared of FFT bins: |X[k]|2, SIR = 8 dB, Pc/Pn =5dB. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100

    43 Performance comparison of MB-OFDM receiver assuming ideal

    NBI detection. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102

    44 The structure of proposed NBI suppression scheme. . . . . . . . . . . 103

    45 The amplitude and step response of adaptive analog notch filter,

    Anotch = −20dB, fBW = 24MHz. . . . . . . . . . . . . . . . . . . . . 104

    46 Simulated BLER versus fBW of AANF. . . . . . . . . . . . . . . . . 105

    47 Notch filter based on band-pass feed-forward cancelation. . . . . . . . 106

    48 Circuit structure of notch filter with a single notching frequency,f0controlled by the digital word, W . . . . . . . . . . . . . . . . . . . . 106

    49 AANF frequency response versus W , Anotch = −40dB. . . . . . . . . 108

    50 AANF filter characteristics versus W , Anotch = −40dB. . . . . . . . . 109

    51 The working flow of proposed NBI suppression scheme. . . . . . . . . 110

    52 Time and frequency domain signals after AGC: without or with

    adaptive analog notch filter, SIR=−10dB, fint = 20, Anotch =−20dB, fBW = 24MHz. . . . . . . . . . . . . . . . . . . . . . . . . . 112

  • xiv

    FIGURE Page

    53 Performance comparison of MB-OFDM receiver, ideal NBI detec-

    tion, 200Mbps date rate, UWB channel model 1. . . . . . . . . . . . 114

  • 1

    CHAPTER I

    INTRODUCTION

    A. Background

    Thanks to its convenience, wireless communication is a promising technology and is

    becoming dominant among all last-one-mile technologies. To transmit data symbols

    over the air efficiently, wireless transceivers often involve carrier modulation. In the

    transmitter, after digital-to-analog (D/A) conversion, the complex baseband signals

    are modulated to high frequency signals by I (in phase) and Q (quadrature) mod-

    ulator. At the receiver, the received signals are converted into the baseband by IQ

    demodulator. Then some optimal or sub-optimal detectors can be used for data re-

    covery. Due to the RF imperfections and the uncertainty of propagation channel, the

    received signals could be seriously distorted and interfered by other wireless systems.

    As shown in Fig. 1, to compensate various distortions and interferences, the

    digital receiver requires efficient carrier and timing recovery, channel estimation and

    interference rejection schemes. This family of algorithms is generally referred to as

    signal acquisition. Efficient and robust signal acquisition algorithms are very critical

    for the success of data recovery. Note that, for some wireless systems, only some

    blocks in Fig. 1 are required and their processing flow may be different from the order

    shown in Fig. 1.

    References [1] - [4] are excellent books on carrier and timing recovery that have

    been reported in the literature. In addition, most of the existing results on interference

    rejection were well summarized in [5]- [6] and their references. The goal of this

    dissertation is to build a general framework for signal acquisition and explore high-

    The journal model is IEEE Transactions on Automatic Control.

  • 2

    Digital Receiver

    ModulatorIQ

    D/A

    ChannelPropagation

    TransmitterBaseband

    IQ DemodulatorA/D

    InterferenceRejection

    TimingRecovery

    CarrierRecovery Detector

    Data

    ChannelEstimation

    Fig. 1. Signal acquisition at digital receiver.

    performance algorithms for various wireless systems. Note that channel estimation is

    not the focus of this dissertation and some existing work on channel estimation can

    be found in [1]- [2].

    Depending on whether the receiver knows the transmitted data or not, the signal

    acquisition schemes can be classified into three broad categories: data-aided (DA),

    decision-directed (DD), and non-data aided (NDA) or blind. All these schemes will be

    addressed in this dissertation, e.g., the DA scheme is used for coarse synchronization

    of OFDM receivers (Chapter IV), the DD scheme is used for fine synchronization of

    OFDM receivers (Chapter V), and NDA schemes are discussed in Chapters II and

    III.

  • 3

    1. Timing Recovery

    In digital communication receivers, the received signal is sampled by means of an

    analog to digital (A/D) converter. However, there is an arbitrary symbol timing delay

    between the receiver and transmitter, which is not known by the A/D converter. To

    get correct data strobe for detection, the digital receiver requires timing recovery. At

    first, the timing offset is estimated using the received data and then a timing error

    controller is used to compensate the estimated timing offset. Depending on the type

    of compensation method used, timing recovery schemes can be classified into two

    classes: feedback and feedforward. In the feedback timing recovery scheme shown

    Data Detector

    Free RunningClock

    A/D InterpolatorDecimator

    Timing OffsetEstimator

    Loop filter

    To

    Fig. 2. Feedback timing recovery scheme.

    in Fig. 2, the estimated timing error is fed back to adjust the time offset of digital

    interpolator (or A/D converter) and the resulting new samples are used to estimate

    the residual timing error. Based on this iterative operation, the system timing error

    converges to the stable operation point. Thanks to its simplicity, the feedback scheme

    is often used in practice. However, for certain initial values of the timing error, the

    acquisition time of blind feedback timing recovery schemes could be extremely large.

    This problem is called hang-up [7]- [9] and will be addressed in Chapter II. As shown in

    Fig. 3, a typical feedforward timing recovery scheme uses the estimated timing offset

  • 4

    To

    Free RunningClock

    A/D

    Timing OffsetEstimator

    InterpolatorDecimator Data Detector

    Fig. 3. Feedforward timing recovery scheme.

    to control the digital interpolator. Since there is no feedback between the timing

    offset estimator and timing error controller, the hang-up phenomenon is avoided in

    feedforward systems [10]- [15]. However, the feedforward timing estimator presents

    the self-noise which is embedded in most of blind estimators [16].

    2. Carrier Recovery

    Due to the imperfections in oscillators and Doppler shifts introduced by the propaga-

    tion channel, carrier frequency and phase offsets are present in the received signal. The

    carrier synchronization task resumes to estimating and compensating these offsets at

    the receiver. Compared to single carrier systems, orthogonal frequency division mul-

    tiplexing (OFDM) systems [17] are very sensitive to carrier frequency offsets. Several

    schemes have been proposed for coarse estimation of the carrier frequency offset [18]-

    [26], which are not efficient in terms of performance and complexity. Therefore, an

    important effort in this dissertation is allocated to the problem of carrier synchroniza-

    tion for OFDM systems. For OFDM receivers, the performance of carrier frequency

    offset estimator depends on the initial timing synchronization. Therefore, the timing

    synchronization of OFDM receivers will be also explored in this work.

    As shown in Fig. 4, the synchronization of OFDM systems consists of two steps:

    coarse and fine synchronization. The coarse synchronization step assumes the task of

  • 5

    removing the large timing and carrier frequency offsets before the fast Fourier trans-

    form (FFT) block. After coarse synchronization, the received signal is converted into

    the frequency domain and might still present residual phase and timing errors. With-

    out compensation, these residual errors could introduce significant performance loss.

    Since these residual errors are small in magnitude, fine synchronization is normally

    performed in the frequency domain using tracking.

    Coarse Synchronization FFT Synchronization

    Fine

    Fig. 4. Coarse and fine synchronization of OFDM receiver.

    3. Interference Mitigation

    Ultra-wideband (UWB) [27] systems provide high data rate, low-power, huge spatial

    capacity and high precision ranging for the quickly growing home networking market.

    Since UWB systems use very low transmission power and operate at license-free

    bands [28], a narrowband interference (NBI) might represent a very critical obstruc-

    tion for the successful operation of UWB systems. It is found that the NBI could cause

    serious performance loss for these systems. Conventional NBI mitigation schemes [5]-

    [6] requires high precision ADC and are not feasible for UWB receivers. Designing

    low complexity and robust NBI mitigation schemes for UWB receivers will represent

    an additional major research thrust of this work.

    B. Organization of Dissertation

    The unifying feature of this work is the development of robust and efficient signal

    acquisition and interference cancellation schemes for general narrowband, wideband

  • 6

    and ultra wideband communication systems. Novel and high performance signal

    synchronization schemes are first proposed for narrowband communication systems

    in Chapters II and III of this work. In Chapters IV and V, the analysis is focused on

    the important problem of designing robust and efficient signal acquisition algorithms

    for the class of wideband systems belonging to the OFDM family. Finally, Chapter VI

    explores the applicability and robustness of the wideband signal interception schemes

    to ultra wideband systems. Also, Chapter VI proposes a novel, robust and high

    performance narrowband interference mitigation scheme for the family of multiband

    OFDM-based UWB transceivers.

    All the proposed signal acquisition and interference mitigation schemes are in-

    troduced gradually starting with narrowband systems, then extending the analysis

    to wideband and ultra wideband systems. Fig. 5 illustrates how this dissertation is

    structured.

    Narrowband

    carrier

    modulated

    systems

    Single

    modulated

    systems

    carrier

    frequencyHighly

    selective

    selectiveFrequency

    Flat−fadingAWGN or

    Coarse synchronization (time domain)

    Fine synchronization(frequency domain)

    Narrowbandinterference mitigation

    Chapter III(feedforward)

    Timing Recovery

    (feedback)Timing Recovery

    Chapter II

    Chapter IV

    Chapter VI

    Chapter V

    Ultra−wideband

    System bandwidth Propagation channel

    WidebandMultiple

    Fig. 5. Content and organization of dissertation.

    In Chapters II and III, we will focus on exploring the high-performance timing

  • 7

    recovery schemes for narrowband systems, where the wireless propagation channel

    can be assumed to be AWGN channel or flat fading channel.

    Chapter II deals with the well known hang-up problem for blind feedback timing

    recovery schemes. When the initial timing offset of blind feedback timing recovery

    scheme is close to half of symbol period, the acquisition time becomes extremely long.

    A novel low complexity anti-hangup algorithm will be proposed to resolve the hang-

    up problem. Simulation results show that the proposed algorithm greatly reduces the

    probability of hang-up and speeds up the feedback timing recovery.

    The performance of blind feedforward timing recovery scheme will be analyzed in

    Chapter III. It is found that the feedforward timing recovery schemes exhibit a large

    self-noise especially for systems with small roll-off factors. The statistical analysis

    of self-noise leads to a novel prefiltering scheme, which can be used to reduce the

    self-noise of blind feedforward timing estimators.

    Chapter IV, V and VI deal with high performance synchronization and inter-

    ference mitigation algorithms for wideband and ultra-wideband systems, where the

    propagation channel could be highly frequency-selective.

    Chapter IV deals with coarse carrier and timing synchronization for OFDM re-

    ceivers. A novel and robust synchronization scheme is proposed for efficient carrier

    frequency offset and timing acquisition.

    Fine synchronization (tracking) in frequency domain is discussed in Chapter V.

    A new decision directed algorithm is proposed to track the residual phase and timing

    error. Both theoretical analysis and computer simulations indicate that the pro-

    posed algorithms greatly improve the efficiency and performance of standard OFDM

    receivers.

    Chapter VI is dedicated to the problem of narrowband interference mitigation

    in multiband OFDM (MB-OFDM) UWB systems. A new NBI cancellation scheme,

  • 8

    based on a digital NBI detector and adaptive analog notch filter bank, is proposed to

    reduce the effects of NBI on UWB systems. Simulation results show that the proposed

    scheme improves significantly the performance of UWB receivers, an improvement

    which translates into an SNR gain of 9 dB. It should be noted that some research

    results on the adaptive analog notch filtering were contributed in collaboration with

    other researchers: Burak Kelleci, Timothy W. Fischer and Dr. Aydın İlker Karsilayan.

    In fact, their contributions proved to be of fundamental importance for this study,

    and of very substantial value.

    Some concluding remarks are provided in Chapter VII. Appendices A-F present

    some mathematical results that are used in the previous chapters.

  • 9

    CHAPTER II

    FAST FEEDBACK TIMING RECOVERY FOR NARROWBAND SYSTEMS

    A. Introduction

    Thanks to its simplicity and robust tracking capability, the digital phase-lock loop

    (PLL) is often used for timing recovery [4]. Normally, the digital PLL for timing

    recovery consists of three elements, which include a timing error detector to estimate

    the current timing error, a lowpass loop filter to decrease the estimation error and a

    timing corrector to control the timing error.

    Fast timing acquisition and good tracking performance are desirable but are

    difficult to achieve simultaneously because the design of PLL is a trade-off between

    acquisition time and tracking error. In noise-free channels, as long as the loop filter

    is determined, the average acquisition time should be a constant. However, for some

    timing error detectors [29]- [31], the PLL occasionally presents a very long acquisition

    time, a phenomenon which is referred to as the hangup problem [7]- [9]. The slow

    timing recovery is undesirable especially for burst transmission systems.

    By introducing a hysteresis effect using a special preamble sequence, the au-

    thors of [32] proposed a fast timing recovery scheme to avoid the hangup in the

    decision-directed timing synchronizer [29]. However, this method works only for

    partial-response signals. The hangup problem also exists in blind timing synchro-

    nizers [30]- [31] for linear and non-linear modulations and it requires a more general

    approach.

    To resolve this problem, in this chapter we propose a novel two-step hangup

    robust timing recovery scheme. Based on an initial estimate obtained in the first

    step, we try to reduce the timing uncertainty from [−T/2, T/2] to [−T/4, T/4] (T is

  • 10

    the symbol period), and thus to avoid the hangup in the second step. The increased

    complexity is very low since the proposed scheme can use the same timing error

    detector as that used in the conventional scheme. We provide simulation results for

    two well-known timing error detectors [30]- [31] and show that the proposed scheme

    greatly speeds up the timing recovery for both linearly and nonlinearly modulated

    transmissions [1].

    B. Signal Model and Feedback Timing Recovery

    In AWGN channels, the received signal can be expressed by

    r(t) = x(t) + n(t) , (2.1)

    where x(t) represents the transmitted linearly or non-linearly modulated signal and

    n(t) denotes the band-limited additive noise. In (2.1), we omit carrier frequency offset

    since timing synchronizer to be discussed is robust to such small frequency offset.

    After passing through matched filter, the received signal becomes y(t). At times

    lT − ǫT , the sampled signal can be expressed as

    yl(ǫ) = y(lT − ǫT ) , (2.2)

    where T , l and ǫT represent the data symbol period, the data symbol index and the

    fractional timing offset in the analog to digital converter (ADC), respectively. For

    simplicity, we omit the carrier frequency offset and phase offset in (2.2) since the

    timing estimators we will use later are robust to these offsets.

    The standard digital phase-lock loop (PLL) is often used for timing recovery.

    The digital PLL for timing recovery is usually made up of three elements [1]- [2]: a

    timing error detector (TED), a loop filter and a timing corrector. First, the TED

  • 11

    outputs a symbol-rate error signal e(l), which depends on the time estimates {ǫ̂l}.

    Reference [30] provides the derivation of a lot of timing error detectors which can

    be blind, decision directed or data-aided. In this chapter, we focus on blind timing

    error detectors. Note that the PN sequence synchronization for the spread spectrum

    receiver is out of the interests of this chapter. For linearly modulated signals, the

    Gardner’s TED [30] can be used

    e(l) = Re{

    y∗l−1/2(ǫ̂l−1) [yl−1(ǫ̂l−1) − yl(ǫ̂l)]}

    . (2.3)

    For minimum shift keying (MSK) type signals, the Mengali’s TED [31] and [1] can

    be used

    e(l) := e(l, D) = (−1)D+1Re{

    y2l−1/4(ǫ̂l−1)y∗2l−D−1/4(ǫ̂l−D−1)

    −y2l+1/4(ǫ̂l)y∗2l−D+1/4(ǫ̂l−D)}

    (2.4)

    In (2.4), D is a design parameter and should be a positive integer. It is pointed out

    in [1] that D = 1 is preferable for MSK and D = 2 leads to the best performance for

    Gaussian MSK (GMSK). From the above equation, the sampling clock for Gardner’s

    TED and Mengali’ TED should be at least Ts = T/2 and Ts = T/4, respectively. The

    error signal is used to recursively update the timing estimates in the loop filter. For

    the first-order loop filter, the updating equation is given by:

    ǫ̂l = ǫ̂l−1 − γe(l) , (2.5)

    where γ denotes the step size.

    Finally, the timing estimates ǫ̂l are exploited to control the timing corrector. De-

    pending on the sampling scheme, the timing correction can be realized with different

    methods. If synchronous sampling is used, the timing estimates are fed into a number

  • 12

    controlled oscillator (NCO) to adjust the sampling clock in the ADC. If asynchronous

    sampling is used, as shown in Fig. 6, the sampling clock in ADC is free running and

    the timing errors are corrected by a digital timing interpolator [33]- [34]. Since the

    asynchronous sampling easily leads to high performance all-digital receivers, we will

    focus on asynchronous sampling hereafter. It should be also mentioned that the pro-

    posed scheme works for synchronous sampling. In Fig. 6, after timing acquisition is

    obtained, the decimator outputs symbol-rate samples to data detector. For linearly

    modulated systems, the lowpass filter can be a matched filter while for nonlinearly

    modulated systems, an anti-alias lowpass filter should be used.

    ADC Interpolator

    Loop Filter

    Decimator

    FilterMatched Timing Error

    Detector

    To

    Detector

    ClockFree−running

    r(t) y(t) yl(ǫl)

    ǫ̂l

    yl(ǫ̂)

    e(l)

    Fig. 6. Conventional timing recovery scheme.

    C. Hangup in Digital PLL

    The design of a loop filter is in general a trade-off between acquisition timing and

    tracking performance. In particular, for the first-order loop filter, if the step-size γ is

    reduced, the PLL needs a longer acquisition time, while the tracking error becomes

    smaller.

    The S-curve of the timing error detector is often used to investigate the per-

    formance of PLL and is defined by the conditional expectation of the timing error

  • 13

    signal:

    S(δ) = E{e(l)|δ} , (2.6)

    where δ is the difference between the true time offset and the current timing estimate

    δ = (ǫ̂l−ǫl)T . It is the S-curve that determinates the strength of pulling the PLL from

    the initial timing offset to the stable point δ = 0. Also, the acquisition and tracking

    performance depends on the loop filter. From [1], the loop bandwidth during tracking

    is given by

    BLT =γA

    4 − 2γA , (2.7)

    where A is the slope of S-curve at δ = 0 and γA < 2 is assumed. In noise-free channels,

    the acquisition time of the digital PLL can be approximated by (pp. 403-404, [2])

    Tacq ≈2|ǫ0|BL

    , (2.8)

    which depends on the loop bandwidth and initial timing offset ǫ0. Assuming ǫ′0s are

    uniformly distributed over [−1/2, 1/2], we obtain the average acquisition time

    T̄acq =1

    2BL=

    (2 − γA)TA

    . (2.9)

    Therefore, in noise-free channels, the acquisition time of the digital PLL is often

    assumed to be inversely proportional to the loop bandwidth BL. However, for some

    timing recovery schemes, the digital PLL occasionally requires a very long acquisition

    time. This phenomenon is called hangup and often happens in PLLs [7]- [9].

    In Fig. 7, we plot the S-curves of different timing error detectors in noise-free

    channels. For Gardner’s TED, we assume that the system uses a QPSK modulation

    and a raised cosine (RC) filter with roll-off factor ρ = 0.5. For Mengali’s TED, a

    GMSK modulated system with pre-modulation bandwidth BT = 0.3 is assumed.

    These simulation parameters will be used throughout this chapter.

  • 14

    −0.5 0 0.5−0.8

    −0.6

    −0.4

    −0.2

    0

    0.2

    0.4

    0.6

    0.8

    S−

    curv

    es

    Gardner’s TED, ρ=0.5Mengali’s TED, BT=0.3

    δ/T

    Fig. 7. The S-curves of timing error detectors.

    The above S-curves have nulls at δ = 0, which corresponds to the desirable stable

    point. However, the S-curves have additional zeros at δ = ±T/2. Therefore, if the

    initial timing offset is close to ±T/2, the average output of TED will be very small.

    Due to the small pull-in strength at δ = ±T/2, it often takes the PLL a very long

    time to converge to the stable point.

    In Fig. 8, the Gardner’ TED is used and we plot the time errors versus the number

    of symbols for different initial timing offsets ǫ0. In the case of ǫ0T = 0.45T , the PLL

    presents monotonic convergence. However, when ǫ0T = 0.5T , the PLL fluctuates

    around 0.5T for a long time before it begins to converge toward the stable point.

    D. Anti-hangup Timing Recovery Scheme

    To remedy the hangup problem, one may use different loop filters during acquisition

    and tracking (pp. 405-406, [2]), e.g., to obtain fast acquisition, one may use a larger

    step size, which is replaced by a smaller step size to reduce the mean-square error

    during tracking. It is found that the probability of hangup is greatly reduced (but

  • 15

    0 100 200 300 400 500 600 700 800 900−0.1

    0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    The number of symbols

    δ/T

    BLT=0.005

    ε0=0.5T

    ε0=0.45T

    Fig. 8. Timing errors versus the number of symbols for Gardner’s TED.

    not avoided) by this method.

    From the discussion in Section III, we know that the hangup will happen only

    if the initial timing offset is close to ±T/2. For discussion purpose, we divide the

    whole range of timing uncertainty into two subsets: t1 = [−T/4, T/4] and t2 =

    (T/4, T/2] ∪ [−T/2,−T/4). If the subset of initial timing error is known in the

    receiver, we may be able to avoid hangup by proper controlling. In the timing error

    detector, shifting the samples by T/4 in (2.6), we obtain the shifted S-curves:

    Ss(δ) = E{e(l − 1/4)|δ} . (2.10)

    As shown in Fig. 9, the shifted S-curves are even functions and cross the zero axis at

    δ = ±T/4. Also, Ss(ǫ0) can be estimated as follows

    S1 =1

    L

    L−1∑

    l=0

    e(l − 1/4) , (2.11)

    where we assume that the timing corrector and loop filter are not working during

  • 16

    −0.5 0 0.5−0.8

    −0.6

    −0.4

    −0.2

    0

    0.2

    0.4

    0.6

    0.8

    Ss(

    δ)

    Gardner’s TEDMengali’s TED

    δ/T

    Fig. 9. The shifted version of S-curves.

    t < LT and ǫ̂l = 0 for l < L. Thus, the subset of initial timing offset may be easily

    detected as follows

    ǫ0T ∈ t1 if S1 ≥ 0 or ǫ0T ∈ t2 if S1 < 0 . (2.12)

    If we detect ǫ0T ∈ t1, e(l) should be used to update the timing estimates in the

    loop filter t > LT and the decimator outputs yl(ǫ̂l) for data detection. However, if

    ǫ0T ∈ t2 is detected, to avoid hangup, we should use e(l − 1/2) to control the loop

    filter. Assuming the coarse detections are always correct, we obtain the new S-curves

    in Fig. 10. Besides the stable point at δ = 0, there are two additional stable points

    at δ = ±T/2. If ǫ0T ∈ t1, the PLL will converge to δ = 0. However, if ǫ0T ∈ t2, it

    will converge to δ = ±T/2. Thus, to obtain correct data detection, yl±1/2(ǫ̂l) should

    be used in the decimator for ǫ0 ∈ t2.

    Based on above results, in Fig. 11, we propose a new fast timing recovery scheme

    which assumes a two-step operation:

  • 17

    −0.5 0 0.5−0.8

    −0.6

    −0.4

    −0.2

    0

    0.2

    0.4

    0.6

    0.8

    δ/T

    New

    S−

    curv

    es

    Gardner’s TEDMengali’s TED

    Fig. 10. The new S-curves.

    ADC Interpolator

    Decimator

    FilterMatched Anti−hangup

    Controllor

    Loop Filter

    Timing ErrorDetector

    To

    Detector

    ClockFree−running

    r(t) y(t) yl(ǫl)

    ǫ̂l

    e(l)

    e(l − 1/4)

    l ≤ L

    l > L

    Fig. 11. Fast timing recovery scheme.

  • 18

    1. Feedforward estimation during t ≤ LT according to (2.11). The output of tim-

    ing error detector is switched to the accumulator in the anti-hangup controller,

    then the feedfoward estimation is used to obtain the coarse information (t1 or

    t2) of ǫ0T using (2.12).

    2. Feedback timing recovery during t > LT . Depending on the coarse timing

    information obtained in the first step, the TED selects a proper set of samples

    to estimate timing errors and update the loop filter for tracking.

    The performance of the new scheme depends on the design parameter L. If L

    is too small, the estimate S1 will be unreliable and will result in incorrect timing

    detection, which may further lead to hangup. Assuming hangup will happen only if

    δ0 = ±T/2, the conditional probability

    Ph := Prob {[S1 > 0|ǫ0T = ±T/2] ∪ [S1 < 0|ǫ0T = 0]}

    should be kept very low to avoid hangup. On the other hand, if in the first step we

    choose L too large, the overall acquisition time may be increased compared to the

    conventional scheme. Therefore, small L should be chosen as long as Ph is kept low.

    Although the variance of e(l) in Gardner’s TED can be calculated as [35], it is found

    that e(l) can not be approximated to be Gaussian distributed for nonzero δ. Thus, it

    is very difficult to obtain the theoretical value of Ph. In Fig. 12, we plot the simulation

    results of Ph versus different L′s for both Gardner’s TED and Mengali’s TED. For

    Gardner’s TED with ρ = 0.5, L = 20 may be enough to keep Ph low. However, due to

    its higher order (4th order) nonlinearity (2.4), Mengali’s TED does not perform well

    at low SNR and requires a longer L. As pointed out in [36], averaging e(l, D) over

    different values of D can improve the performance of feedforward estimation. Thus,

  • 19

    0 2 4 6 8 1010

    −4

    10−3

    10−2

    10−1

    100

    Es/N

    o (dB)

    Ph

    L=20L=40

    ρ=0.5, QPSK

    (a) Gardner’s TED

    4 6 8 10 12 14 16 1810

    −4

    10−3

    10−2

    10−1

    100

    Es/N

    o (dB)

    Ph

    L=20, D=2L=20, CombiningL=40, D=2

    GMSK, BT=0.3

    (b) Mengali’s TED

    Fig. 12. Ph versus different L.

  • 20

    in the first step operation, the following estimator can be used to reduce Ph:

    S2 =1

    3L

    3∑

    D=1

    L−1∑

    l=0

    e(l − 1/4, D) . (2.13)

    As shown in Fig. 12 (b), for the same L, the simulated Ph with the above combining

    estimator is much lower than the Ph obtained with estimator (2.4).

    To verify the advantage of the proposed timing recovery scheme, we run simula-

    tions to compare the performance of the different timing recovery schemes. In simula-

    tions, we assume a system with QPSK modulation and a RC filter with ρ = 0.5. For

    conventional and proposed schemes, the step size of loop filter is fixed to γ = 0.013,

    which leads to a loop bandwidth during tracking: BLT = 0.005. Also, L = 20 symbols

    are used during the first step of the proposed scheme.

    For the adaptive loop filter scheme, to speed up the acquisition, a larger step

    size γ = 0.026 is used during initial acquisition. According to (pp. 405-406, [2]),

    for Gardner’s TED, e(l − 1/4) can be used to detect if PLL is in lock. From [30],

    denoting H(f) as the Fourier transform of RC filter, we can express the expectation

    of e(l − 1/4) by

    E[e(l − 1/4)] = cos(2πδ) · 2T

    ∫ 1/2T

    −1/2TH(f − 1

    2T)H(f +

    1

    2T) cos(πfT )df , (2.14)

    which is maximum when the PLL is in lock. A low-pass filter is used to average

    e(l − 1/4)

    Ld(l) = γfLd(l − 1) + (1 − γf )e(l − 1/4) , (2.15)

    where γf denotes the forgetting factor and our simulations assume γf = 0.95. Ld(l)

    is then compared with an appropriate threshold Thr for lock detection. As long as

    the PLL is in lock, the step size of loop filter will be changed to γ = 0.013 to reduce

    the tracking error. A threshold of Thr = 0.11, which is a half of (2.14) when δ = 0

  • 21

    and ρ = 0.5, is used in simulations.

    0 100 200 300 400 500 600 700 800 900

    −0.5

    −0.4

    −0.3

    −0.2

    −0.1

    0

    0.1

    0.2

    0.3

    0.4

    0.5

    The number of symbols

    δ/T

    (a) Conventional timing recovery scheme

    γ=0.013

    0 100 200 300 400 500 600 700 800 900

    −0.5

    −0.4

    −0.3

    −0.2

    −0.1

    0

    0.1

    0.2

    0.3

    0.4

    0.5

    The number of symbols

    δ/T

    (b) New timing recovery scheme

    γ=0.013

    Step 1

    Step 2

    Fig. 13. Performance comparison in AWGN channels, Es/No = 15 dB, 100 runs.

    Fig. 13 shows 100 runs in AWGN channels, using different initial timing offsets

    and noise sequences. Figs. 13 (a), (b) are the simulation results of the conventional

    scheme, the proposed scheme, and adaptive loop filter scheme, respectively. Since all

    schemes use the same loop bandwidth during tracking, they present similar tracking

  • 22

    error. However, their acquisition performance is very different. For the conventional

    scheme, although all runs finally converge to δ = 0, sometimes the acquisition is very

    slow. In 5 out of 100 runs, the convergence to δ = 0 is not achieved even after 300

    symbols. On the other hand, the proposed scheme exhibits fast time recovery. For

    all experiments using the proposed scheme, the convergence is obtained in less than

    150 symbols.

    Interesting enough, the timing acquisition of the proposed scheme is even faster

    than that of conventional scheme when ǫ0T 6= ±T/2. This can be explained by the

    smaller timing uncertainty obtained in the feedfoward estimation of proposed scheme.

    In noise-free channels, assuming ǫ0 are uniformly distributed over [−T/4, T/4] after

    feedforward estimation, we obtain the average acquisition time of proposed scheme

    T̄acq =1

    4BL+ LT , (2.16)

    which is only a half of (2.9) if LT is omitted.

    It is noticed that the proposed scheme presents three stable points (0,±T/2),

    which verifies our previous discussion. However, correct data strobes can be obtained

    by properly controlling the decimator in Fig. 11.

    The performance of conventional and proposed timing recovery schemes is also

    evaluated in flat fading channels. Due to the channel fading, the received signal

    exhibits a large dynamic range, which may lead to a variable loop bandwidth. To

    keep the loop bandwidth constant, before the signal enters the timing loop, a AGC is

    used to normalize the received signal by using the root mean-square of the received

    signal vector. As shown in Fig. 14, compared to the conventional scheme, the proposed

    scheme also presents faster acquisition in fading channels.

    Similar improvements can be found for the timing recovery schemes that assume

    Mengali’s TEDs in MSK-type systems. These results verify that the proposed scheme

  • 23

    0 50 100 150 200 250 300 350 400 450−0.6

    −0.4

    −0.2

    0

    0.2

    0.4

    0.6(a) Conventional timing recovery scheme

    The number of symbols

    δ/T

    γ=0.013

    0 50 100 150 200 250 300 350 400 450−0.6

    −0.4

    −0.2

    0

    0.2

    0.4

    0.6(b) New timing recovery scheme

    The number of symbols

    δ/T

    γ=0.013

    Fig. 14. Performance comparison in flat fading channels, BLT = 0.01, Es/No = 20

    dB, fdT = 0.0002.

  • 24

    greatly speeds up the timing acquisition.

    We remark that it is not free to achieve fast time recovery. Although the same

    TED can be used in the two schemes, the proposed scheme requires an additional

    anti-hangup controller. Thus, the proposed scheme slightly increases the complexity

    of timing synchronization. Furthermore, the original Gardner’s TED assumes only

    two samples per symbol while the proposed scheme requires four samples per symbol.

    It should be mentioned we do not increase the sampling requirement for Mengali’s

    TED which originally assumes four samples per symbol [31].

    E. Summary

    In this chapter we have proposed a new anti-hangup timing recovery scheme which

    assumes a two-step operation. Based on the initial timing estimate obtained in the

    first step, we reduce the timing uncertainty from [−T/2, T/2] to [−T/4, T/4], and

    thus avoid the hangup phenomenon in the second step. For two well-known timing

    error detectors, we show through simulation results that this simple scheme greatly

    speeds up the timing recovery for both linearly and nonlinearly modulated systems.

    Blind feedforward timing estimator [10] was proposed to avoid hang-up problem.

    However, some blind feedforward estimator present error floors at hight SNR, which

    is caused by the self-noise [7]. In next chapter, a novel prefilter will be proposed to

    reduce the self-noise of feedward timing estimators.

  • 25

    CHAPTER III

    JITTER-FREE FEEDFORWARD TIMING RECOVERY FOR NARROWBAND

    SYSTEMS

    A. Jitter-Free Prefilter for Feedforward Timing Recovery of Linear Modulations

    1. Introduction

    Nondata-aided (blind) feedforward timing recovery schemes [10]- [15] enable fast and

    reliable synchronization, and therefore, they have found use in burst transmission sys-

    tems. A general cyclostationary framework that exploits second-order nonlinearities

    to design digital blind feedforward synchronizers was proposed in [13].

    For systems with large excess bandwidth, the performance of the synchronizers

    that exploit the spectral line generated by a nonlinearity is asymptotically (large

    sample) very close to the Cramer-Rao bound. If the excess bandwidth is small,

    significant jitter is induced by data pattern which degrades the performance of the

    timing recovery scheme in mid and high SNRs [16].

    Franks and Bubrouski found that the analog second-order nonlinearity based

    timing synchronizer can be jitter-free if an appropriate prefilter is used [38]. Ref-

    erence [39] extended Franks and Bubrouski result to analog synchronizers that as-

    sume arbitrary nonlinearities. For digital timing recovery, [40]- [41] exploited similar

    prefilters to improve the performance of Gardner’s nondata aided timing recovery

    scheme [30]. By means of simulation results, [42] reported that a similar prefilter

    could be used to improve the performance of the four samples per symbol based feed-

    forward scheme [10]. However, no rigorous theoretical analysis was conducted in [42]

    to justify the jitter-free timing recovery condition.

    In this section, we derive a closed-form expression for the power of self-noise

  • 26

    present in digital blind feedforward timing estimator [15], which is a modified version

    of [14] and needs two samples per symbol. This rigorous derivation of the self-noise

    power is later exploited for designing a prefilter that ensures nearly jitter-free timing

    recovery scheme. The proposed analysis and design can be easily applied to feedfor-

    ward timing recovery schemes that assume oversampling factors larger than or equal

    to three (see e.g., [10]). In essence, it is shown that if the Franks-Bubrouski condition

    is fulfilled, i.e., the frequency response of the equivalent pulse is symmetric with re-

    spect to half the symbol rate and has bandwidth less than symbol rate, then a nearly

    jitter-free digital feedforward timing recovery scheme is obtained for any oversam-

    pling factor. Finally, we find the equivalence between the digital feedforward timing

    estimator and the analog synchronizer can directly lead to the prefilter by utilizing

    the previous results in [39].

    2. Signal Model and Symbol Timing Estimators

    We consider the baseband representation of a linearly modulated signal transmitted

    through an AWGN channel. The receiver input is expressed as

    r(t) =∑

    l

    alhT (t − lT − ǫT ) + w(t), (3.1)

    where al stand for zero-mean unit variance (E|al|2 = 1) independently and identically

    distributed (i.i.d) complex valued symbols with the fourth-order moment E[|al|4] = γ,

    hT (t) is the transmitter’s filter, and w(t) is complex white Gaussian noise with two-

    sided power spectral density N0/2 per component. In (1), ǫ denotes the unknown

    symbol timing delay (normalized by the symbol period T ).

    To simplify the analysis, we assume that the frequency offset has been compen-

    sated before the timing recovery task is performed (see e.g., [13] for such frequency

    compensation schemes). Since the proposed timing delay estimator is insensitive to

  • 27

    carrier phase offsets, we also omit the presence of carrier phase offset in (3.1). After

    matched filtering with hR(T ), the resulting signal x(t) is oversampled by Ts := T/P ,

    with the oversampling factor P ≥ 2, and the received sequence is given by

    x[n] =∑

    l

    alhc[n − lP − ǫP ] + v[n], (3.2)

    where x[n] := x(nTs), v[n] := w(t) ⊗ hR(t)|t=nTs , hc[n] := hc(nTs), hc(t) := hT (t −

    ǫT )⊗hR(t), and ⊗ denotes the convolution operator. We assume that hc(t) is a raised

    cosine (RC) pulse of bandwidth [−(1 + ρ)/2T, (1 + ρ)/2T ], with the roll-off factor ρ

    (0 < ρ < 1) and its Fourier transform (FT) is denoted Hc(F ).

    To estimate the timing offset, the second-order cyclostationary statistics of an

    observation vector with N symbols will be exploited. The sample cyclic correlation

    coefficient at cycle k and lag τ (integer) is given by

    R̂x(k, τ) =1

    NP

    NP−τ−1∑

    n=0

    x∗[n]x[n + τ ]e−j2πkn/P , (3.3)

    where ∗ denotes the conjugation operator. Based on (3.3), a general estimator was

    derived by Gini and Giannakis [13] for oversampling factors P ≥ 3

    ǫ̂ = − 12π

    arg

    Lg∑

    τ=−Lg

    1

    G(τ)R̂x(1, |τ |)

    , (3.4)

    For any even and real hc(t), e.g., the RC pulse, G(τ) is a real and even function and

    defined by

    G(τ) :=P

    T

    ∫ 1/2T

    −1/2THc(F −

    1

    2T)Hc(F +

    1

    2T)ej2πτTF/P dF. (3.5)

    The upper limit Lg of summation (3.4) is fixed as the maximum lag τ that

    provides a non-zero value for |G(τ)|(≫ 0). In (3.4), Lg cyclic correlations R̂x(1, τ)

    are averaged with weighting factors G−1(τ). Simulation results in [12] shows that

    estimator (3.4) with small Lg exhibits good performance. Thus, in this chapter we

  • 28

    use O&M estimator (Lg = 0) for P ≥ 3. A new timing estimator using two samples

    per symbol was proposed by Lee in [14]. However, Lee’s estimator is biased for large ρ

    and the authors in [15] modified Lee’s estimator and obtained an unbiased estimator.

    Therefore, we introduce herein the following symbol timing estimators1

    ǫ̂=− 12π

    arg{Γ(ǫ)} , (3.6)

    Γ(ǫ)=

    R̂x(1; 0) − j G(0)G(1)ℜ{R̂x(1; 1)}, P = 2

    R̂x(1; 0), P ≥ 3.

    The performance of the above estimators is very close to modified Cramer-Rao

    bound (MCRB) [1] for large roll-off factors ρ as shown in [15] and [12]. However, for

    small ρ, a large error floor will be caused by self-noise, a fact which can be observed

    next section for oversampling factors P = 2. Similar results can be observed for

    P ≥ 3 and are not presented here for space limitation.

    To reduce this self-noise, we introduce a prefilter hpre(t) with FT Hpre(F ). Since

    this chapter focuses on the compensation of self-noise, we omit the additive noise

    hereafter. Thus, we replace hc[n] in (2) with h[n] := h1(nTs), where h1(t) := hc(t) ⊗

    hpre(t). Also, in (3.5), Hc(F ) is replaced with H1(F ) := Hc(F )Hpre(F ), the FT of

    real-valued filter h1(t), which is assumed to be an even function.

    3. Nearly Jitter-Free Prefilter

    At first, let us consider a special case ǫ = 0. For P = 2, considering R̂x(1; 0) is

    real-valued, to get jitter-free estimation ǫ̂ = 0, we need

    ℜ{R̂x(1; 1)} =1

    2Nℜ{∑

    l

    k

    a∗l ak2N−2∑

    n=0

    h[n − 2l]h[n + 1 − 2k](−1)n}

    (3.7)

    1Notations ℜ{·} and ℑ{·} denote the real and imaginary part, respectively.

  • 29

    to be zero. This can be obtained if we let h[n − 2l]h[n + 1 − 2k] = 0, or equivalently

    the product of two consecutive samples is zero. Since h[2n] = h1(nT ) is the data

    strobe with unit amplitude, it follows that

    h[2n + 1] = h1(nT + T/2)

    =∫ 1/T

    −1/TH1(F )e

    j2πFT (n+1/2)dF

    =∫ 1/T

    0[H1(F ) − H1(F − 1/T )]ej2πFT (n+1/2)dF (3.8)

    should be zero for any n, which results in

    H1(F ) = H1(F − 1/T ) , F ∈ (0, 1/T ) . (3.9)

    Considering H1(F ) is an even function, we find that, for F > 0, H1(F ) should be

    symmetric around 1/2T (or −1/2T , for F < 0) and with bandwidth less than 1/T .

    For RC pulse, the above result leads to the prefilter

    hpre(t) = hc(t)cos(2πt/T ) . (3.10)

    Actually, (3.8) was also used in [40] to derive the prefilter for Gardner’s feedback

    scheme [30].

    Motivated by this special result, we want to check if the above prefilter has

    removed the self-noise of estimator (3.6) for general cases (ǫ 6= 0). However, for

    non-zero timing offset case, the above analysis can not verify the estimator with

    the prefilter (3.10) is jitter-free. Then we resort to calculate the mean-square error

    (MSE) of the estimators in closed form. For simplicity, we will keep to the self-noise

    calculation for P = 2.

    The estimation error of (3.6) can be expressed as

    ǫ̂ − ǫ = − 12π

    arg{Γ(ǫ)ej2πǫ} . (3.11)

  • 30

    Taking the tan-function of both sides of (3.11), we obtain

    tan[2π(ǫ̂ − ǫ)] = −ℑ{Γ(ǫ)ej2πǫ}

    ℜ{Γ(ǫ)ej2πǫ} ≈ 2π(ǫ̂ − ǫ) , (3.12)

    where the last approximation holds whenever the estimation error is small since

    tan(x) ≈ x for |x|

  • 31

    terms involved in (3.16)- (3.17). In Fig. 15, we plot the ratios Gm(1, 0)/Gm(0, 0), m =

    1, 2, for different roll-off factors ρ, where

    Gm(0, V ) :=∫ 1/2T

    −1/2THm2 (F, V )dF ,

    Gm(1, V ) :=∫ 1/2T

    −1/2THm2 (F, V )cos(πFT )dF . (3.18)

    0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.90.92

    0.93

    0.94

    0.95

    0.96

    0.97

    0.98

    0.99

    1

    ρ

    G1(1,0)/G

    1(0,0)

    G2(1,0)/G

    2(0,0)

    Fig. 15. Plots of Gm(1, 0)/Gm(0, 0) for m = 1, 2 versus various roll-off factors ρ.

    Fig. 15 shows that Gm(1, 0)/Gm(0, 0) is very close to 1, especially for small ρ.

    Recalling that G1(1, 0) = G(1) and G1(0, 0) = G(0), we can approximate G(1) ≈

    G(0). Thus, we can approximate (3.16)-(3.17) by

    Q1(V ) ≈ 2G2(0)G1(0,−V )[G1(0, V ) − G1(1, V )], (3.19)

    Q2(V ) ≈ 2G2(0)[G2(0, V ) − G2(1, V )] . (3.20)

    Similarly, we can also approximate Gm(1, V )/ Gm(0, V ) ≈ 1 for small V . Note

    that for large V , sin2(πNV T )/(π2V 2) ≈ 0, we can approximate Q1(V ) ≈ 0 and

    Q2(V ) ≈ 0 in (3.15). Thus, we obtain E[Qss] ≈ 0, which means that we can reduce

  • 32

    the self-noise with the prefilter defined in (3.10).

    0 5 10 15 20 25 30 35 40 4510

    −8

    10−7

    10−6

    10−5

    10−4

    10−3

    10−2

    10−1

    16QAM, ρ=0.25 N=100

    Without prefilter

    With prefilter

    MCRB

    Es/N

    o (dB)

    Mea

    n S

    quar

    e E

    rror

    Fig. 16. Comparison of symbol timing estimators for P = 2.

    To verify this result, we compare the performance of estimators with and without

    prefilter for systems with low roll-off factors (ρ = 0.25) by simulations in Fig. 16. For

    each SNR value, we average the performance of 1, 000 Monte-Carlo runs over different

    timing offsets. At mid and high SNRs where the performance is dominated by the

    self-noise, no obvious error floor is observed for SNR less than 35dB for the estimator

    with prefilter and its performance is closer to the MCRB. At low SNRs where the

    additive noise is dominant, the prefilter does not degrade the performance since the

    same excess bandwidth is preserved. In Fig. 17, we also compare the performance of

    estimator with prefilter for different N . Except for very high SNR, there is still no

    error floor for short observation interval N=24, which verifies that nearly jitter-free

    timing recovery can be obtained by prefilter..

    The above analysis can be easily extended to O&M estimator. By calculating

    self-noise in (3.6) for (P ≥ 3), we find the same filter as (3.10) can be used to remove

  • 33

    0 5 10 15 20 25 30 35 40 4510

    −8

    10−7

    10−6

    10−5

    10−4

    10−3

    10−2

    10−1

    Es/N

    o (dB)

    Mea

    n−S

    quar

    e E

    rror

    N=24, with prefilterN=24, MCRBN=100, with prefilterN=100, MCRB

    16QAM, ρ=0.25, P=2

    Fig. 17. Comparison of the performance of estimator for different N .

    the self-noise for O&M estimator, and this theoretical result can be used to justify the

    simulation results in [42]. Also, we find that this prefilter brings a new explanation

    for (3.4). If we replace hpre(t) by an equivalent FIR hpre[n] of LP +1 taps and assume

    R̂x(1, |τ |) are related to the samples at the input of the prefilter, the timing estimator

    (3.6) for P ≥ 3 can be viewed as

    ǫ̂ = − 12π

    arg

    LP∑

    τ=−LPrhpre [τ ]R̂x(1, |τ |)

    , (3.21)

    where rhpre [τ ] :=∑LP +|τ |

    τ1=|τ | hpre[τ1]h∗pre[τ1 − |τ |] can be viewed as the autocorrelation

    coefficient at lag τ of the impulse response of the prefilter. Thus, to remove the

    self-noise, the 2Lp + 1 cyclic correlations in (3.4) should be averaged with weighting

    factors that depend on the prefilter’s autocorrelation coefficients.

    We remark that in frequency-selective channel the symmetry of H1(F ) can not

    be obtained unless the channel is perfectly known. Thus, jitter free timing recovery

    cannot be obtained by simple prefiltering in this channel.

  • 34

    4. Summary

    We have derived a closed-form expression for the power of self-noise for a blind feed-

    forward symbol timing estimators that assume oversampling factors equal to 2. It

    has been shown that an appropriate prefilter after the receiver’s matched filter can

    be utilized to reduce the jitter. This result can be easily extended for estimators with

    higher (P ≥ 3) oversampling factors. Simulation results prove that the synchronizer

    which assumes such a prefilter is nearly jitter-free even in the presence of a finite

    (reduced) number of samples as long as linear-modulated signals are assumed. Such

    prefilters actually can improve performance of square-law estimators for MSK and

    GMSK-type signals, a topic which will be researched in the next section.

    B. Jitter-Free Prefilter for Feedforward Timing Recovery of GMSK Modulations

    1. Introduction

    According to Laurent’s approximation [43], minimum-shift keying (MSK)-type signals

    can be viewed as a superposition of some linearly modulated waveforms. Based on this

    interpretation, several blind feedforward timing estimators, similar in structure to the

    schemes proposed for linear modulations [10]- [15], were proposed for MSK/Gaussian

    MSK (GMSK) modulations in [44]- [47]. At low signal-to-noise ratios (SNRs), the

    estimators [46]- [47] approach the Cramer-Rao bound (CRB) and outperform the

    timing recovery schemes [44]- [45]. By exploiting jointly the information contained

    in the in-phase and quadrature components of the received signal, a novel timing

    recovery algorithm with improved performance relative to the timing estimator [46] is

    proposed. Unfortunately, due to self-noise (jitter), the performance of timing recovery

    schemes [46] and [47] for GMSK modulations is far from CRB at mid and high SNRs.

    Motivated by the fact that prefilter based structures can improve the performance of

  • 35

    feedforward timing estimators for linear modulations (see [48] and the references cited

    therein), this chapter proposes a similar prefilter to reduce the self-noise of timing

    recovery schemes [46]- [47]. Computer simulations show that the proposed prefilter

    based timing recovery schemes exhibit improved performance at mid and high SNRs

    with respect to the existing schemes [46]- [47].

    2. Signal Model and Estimators

    According to Laurent’s approximation [43], an MSK-type signal can be approximated

    by

    s(t) ≈∑

    l

    exp [jπ

    2

    l∑

    i=1

    Ii]hT (t − lT ) , (3.22)

    where

    hT (t) :=L−1∏

    i=0

    p(t + iT ) , (3.23)

    p(t) :=

    sin[πq(t)], 0 ≤ t ≤ LT

    p(2LT − t), LT < t ≤ 2LT

    0, otherwise.

    (3.24)

    In (3.22)-(3.24), T denotes the symbol period, Il stands for the zero mean and in-

    dependently and identically distributed (i.i.d) binary information sequence and q(t)

    is the phase response of the modulator, supposed of length L. If the signal s(t) is

    assumed to be transmitted through an AWGN channel, after carrier frequency and

    phase offset compensation, the receiver output can be expressed as

    r(t) = s(t − ǫT ) + w(t), (3.25)

    where w(t) stands for complex white Gaussian noise with two-sided power spectral

    density N0/2 and ǫT denotes the unknown symbol timing delay. Since the MSK-

  • 36

    type signal can be viewed as a linearly modulated signal with pulse shaping filter

    hT (t), the received signal is processed through the matched filter hR(t) := hT (−t)

    and oversampled with the sampling period Ts = T/P (P ≥ 3). Consequently, the

    received sequence can be expressed as

    x(n) =∑

    l

    alh(n − lP − ǫP ) + v(n) , (3.26)

    where al ∈ {±j} for odd l, al ∈ {±1} for even l, v(n) := w(t) ⊗ hR(t)|t=nTs, h(n) :=

    hT (t)⊗hR(t)|t=nTs, and ⊗ denotes the convolution operator. Also, let H(f) stand for

    the Fourier transform (FT) of h(n). After performing some approximations (accurate

    only for low SNRs) on the log-likelihood function, Morelli and Vitetta derived the

    blind feedforward timing estimator [47]

    ǫ̂ := − 12π

    arg

    {P−1∑

    τ=0

    |R(τ)|2e−j2πτ/P}

    , (3.27)

    where R(τ) is obtained by an operation of squaring the observation vector of length

    L0:

    R(τ) =L0−1∑

    n=0

    (−1)n+1x2(nP + τ) , (3.28)

    and the modulating factor (−1)n+1 is used to remove the time-varying effects intro-

    duced by the data modulation. The above estimator is very similar to the digital

    timing recovery algorithm proposed for linear modulations in [10].

    A feedforward clock recovery configuration is shown in Fig. 18. The estimated

    value ǫ̂ is post-processed by a lowpass filter with saw-tooth nonlinearity [10] to con-

    trol the interpolator. The prefilter hpre(n) will be later introduced to improve the

    performance of timing estimator by reducing its self-noise.

    Note that (3.26) can be rewritten as

    x(n) =∑

    l

    b2lh(n − 2lP − ǫIP ) + j∑

    l

    b2l+1h(n − 2lP − ǫQP ) + v(n) , (3.29)

  • 37

    FixedClock ~

    Timing Estimator

    Matched

    To

    DetectorInterpolator

    Filter Prefilterx[n]Postprocessor

    r(t)

    Fig. 18. Feedforward clock recovery scheme.

    where bl ∈ {±1}, ǫI = ǫ and ǫQ = ǫ + 1. Thus, an MSK-type signal can be viewed

    as an offset QPSK (OQPSK) modulated signal and both in-phase (I) and quadrature

    (Q) components of (3.29) are cyclostationary with period 2P . Following the steps

    in [13], the cyclic correlations of the real and imaginary components of x(n) are found

    to be

    n

    {ℜ[x(n)]}2e−j2πn/2P ∝ e−jπǫI ,∑

    n

    {ℑ[x(n)]}2e−j2πn/2P ∝ e−jπǫQ , (3.30)

    where ℜ[x] and ℑ[x] denote the real and imaginary parts of the complex-valued num-

    ber x, respectively. In [46], estimates of the timing delay are obtained by exploiting

    separately the in-phase (I) and quadrature (Q) components of the received signal,

    and are given by

    ǫ̂I = −1

    πarg

    L0P−1∑

    n=0

    {ℜ[x(n)]}2e−j2πn/2P

    ,

    ǫ̂Q = −1

    πarg

    L0P−1∑

    n=0

    {ℑ[x(n)]}2e−j2πn/2P

    . (3.31)

    Estimator [46] exploits separately these estimates to obtain an estimate of the timing

  • 38

    delay ǫ. However, we notice that if both I and Q components are jointly exploited

    to estimate ǫ, the performance of [46] can be further improved. Taking (3.30) into

    account, this new estimator takes the form:

    ǫ̂ := − 1π

    arg

    L0P−1∑

    n=0

    {ℜ[x(n)]}2e−j2πn/2P + ejπ ·L0P−1∑

    n=0

    {ℑ[x(n)]}2e−j2πn/2P

    ,

    = − 1π

    arg

    L0P−1∑

    n=0

    ℜ[x2(n)]e−jπn/P

    . (3.32)

    0 5 10 15 20 25 30 35 40 4510

    −7

    10−6

    10−5

    10−4

    10−3

    10−2

    10−1

    WN estimator (I only)WN estimator (Q only)New estimator MV estimatorMCRB

    Nor

    mal

    ized

    Mea

    n S

    quar

    e E

    rror

    Eb/N

    o (dB)

    L=100 GMSK, BT=0.3

    Fig. 19. Performance comparisons of various estimators for GMSK modulations.

    In Fig. 19, computer simulations illustrate the performance of the above men-

    tioned estimators for GMSK modulations. The performance of the new estimator

    (3.32) is always better than that corresponding to the estimator that relies only on

    either the I or Q-component [46]. If no phase offset is assumed, the performance

    of (3.32) is even slightly better than than that exhibited by the Morelli and Vitetta

    (MV) estimator. Unfortunately, (3.32) is sensitive to residual phase offsets while the

    MV estimator works well for large phase offsets. Because the prefilter to be discussed

  • 39

    in next section provides more improvement for GMSK modulations rather than for

    MSK modulations, in what follows we will focus only on designing prefilter based

    timing estimators for GMSK modulations.

    3. Self-Noise Compensating Prefilter

    At low SNRs, the performance of MV estimator is very close to CRB since it is

    an approximate ML estimator at low SNRs. However, the MV estimator is not

    necessarily good at medium and high SNRs since its approximation is not accurate

    anymore. It is known that the self-noise dominates the performance of such estimators

    at high SNRs. The simulation results presented in Fig. 19 show that for GMSK

    modulations the performance of the MV estimator is far away from CRB at high

    SNRs, which indicates that large self-noise exists in (3.27).

    For linear modulations, some prefilters have been found to reduce the self-noise in

    analog synchronizers [38] as well as in digital synchronizers [40] and [48] that exploit

    second-order nonlinearities. Since the MV estimator is very similar to the second-

    order nonlinearity based timing recovery algorithm [10] 2, we expect that a properly

    designed prefilter can also reduce the self-noise of both the MV estimator and the

    new estimator (3.32), which is also similar in structure to [10].

    Both estimators (3.27) and (3.32) exploit two key facts: an MSK-type signal is

    equivalent to an OQPSK modulation, and the symbol period of the I and Q compo-

    nents of an MSK-type signal is 2T . Therefore, we find that the excess bandwidth of

    MSK-type signals is the part of signal bandwidth larger than 1/4T instead of 1/2T

    (which holds for linear modulations). Thus, to reduce the self-noise, we introduce

    2The only difference is that for removing the effect of possible phase offsets in(3.27), the absolute value of R(τ) is computed before evaluating the complex Fouriercoefficients.

  • 40

    a prefilter similar in structure to the one proposed in [48] for feedforward timing

    recovery in linear modulations

    hpre(n) := cos(πn/P )h(n) . (3.33)

    In the frequency domain, (3.33) is just the shifted version (shifted by 1/2T ) of h(n)

    such that the FT of hpre(n) ⊗ h(n), Hpre(f) · H(f), is almost3 symmetrical around

    1/4T .

    0 0.5 1 1.5 20

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.8

    0.9

    1|H(f)||H

    pre(f)|

    |H(f)Hpre

    (f)|

    f/T

    Am

    plitu

    de R

    espo

    nse

    Fig. 20. Amplitude response of filters.

    This prefilter is shown in Fig. 20 for GMSK modulations. This prefilter is ex-

    pected to reduce the self-noise of timing recovery estimators. Similar results have

    been reported and analytically proven in [48] for linear modulations. To investi-

    gate the performance of prefilter based timing recovery estimators, simulation results

    for GMSK modulations with pre-modulator bandwidth BT = 0.3 are illustrated in

    3To maintain Hpre(f)·H(f) strictly symmetrical around 1T , an additional bandpassfilter may be required. However, for GMSK, we find that (3.33) is sufficient to obtainthe desired symmetry.

  • 41

    Fig. 21. The observation length is 100 and the oversampling factor is P = 4. All

    simulation results are obtained by performing 2,000 Monte-Carlo trials averaged over

    different timing offsets for each SNR value. Also, the modified CRB (MCRB) [1] is

    plotted.

    0 5 10 15 20 25 30 35 40 4510

    −7

    10−6

    10−5

    10−4

    10−3

    10−2

    10−1

    MV estimator, without prefilterMV estimator, with prefilterNew estimator, without prefilterNew estimator, with prefilter MCRB

    Eb/N

    o (dB)

    Nor

    mal

    ized

    Mea

    n S

    quar

    e E

    rror

    Fig. 21. The performance of estimators with and without prefilter.

    MCRB(ǫ) =1

    2π2 · L0 · Eb/No ·∫∞−∞ Tg

    2(t)dt, (3.34)

    where g(t) is the frequency response of modulator. The simulation results indicate

    that the proposed prefilter improves significantly the performance of the MV estimator

    and the new estimator (3.32) for SNR values equal to and larger than 15 dB, while

    at low SNRs their performance is almost the same. Also, the proposed prefilter

    can work for MSK modulations (a result that is not detailed herein). Since the

    excess bandwidth of GMSK modulations is much smaller than that corresponding to

    MSK modulations, much more significant self-noise is induced in (3.27) and (3.32).

    Therefore, the usage of self-noise compensating prefilters leads to larger performance

  • 42

    improvement for GMSK modulations. We notice further that there is still an error

    floor at very high SNRs, a fact which can be explained by the Laurent approximation

    error in (3.22) and the approximation made in [48] when deriving the expression of

    self-noise. As a side observation, we remark that the above mentioned improvement

    is obtained by increasing the complexity of the receiver since an additional prefilter

    besides the matched filter HR(t) is required. However, we have found that a prefilter

    with 20 taps is precise enough for GMSK modulations with BT=0.3.

    4. Summary

    A new timing estimator with improved performance relative to the Wu and Ng esti-

    mator [46] has been proposed for MSK-type modulations. The performance of this

    new estimator is slightly better than that of Morelli and Vitetta estimator [47]. How-

    ever, both estimators are sensitive to phase offset errors. We have also derived a

    new digital prefilter to reduce the self-noise for both the MV estimator [47] and the

    WN estimator [46], a result which represents an extension of [48]. Simulation results

    assessed in AWGN channels show that the estimators with the proposed prefilter are

    very close to CRB for practical SNR. For GMSK modulations with pre-modulator

    bandwidth of 0.3, this simple prefilter improves the performance of existing timing

    recovery schemes for SNRs equal to or larger than 15 dB. The proposed prefilter

    based timing recovery estimators are almost jitter free and present slightly increased

    complexity.

  • 43

    CHAPTER IV

    EFFICIENT COARSE FRAME AND CARRIER SYNCHRONIZATION FOR

    OFDM SYSTEMS

    A. Introduction

    In last two chapters, we have discussed timing recovery for single carrier modulated

    systems, where we assume AWGN or flat fading channel. As the system bandwidth

    increases, the propagation channel becomes more frequency selective. Due to its high

    complexity and slow convergence, the conventional equalizer (time domain) is not fit

    for wideband systems. Thanks to the simple frequency domain equalization, orthog-

    onal frequency division multiplex (OFDM) systems have been adopted in wideband

    systems.

    However, due to the sensitivity of OFDM systems to synchronization errors (espe-

    cially to carrier frequency offset), reliable synchronization schemes must be designed

    for these systems. By exploiting the known structure of a training symbol or a cyclic

    prefix, several schemes have been proposed for coarse estimation of the carrier fre-

    quency offset and timing delay [18]- [26], and [49].

    In [19], the training symbol contains two identical halves [+A +A], and the

    timing delay estimator is obtained by searching for the peak of correlation between

    the first and second halves of the received symbols. By comparing the phase difference

    between the identical parts, a coarse frequency offset estimator is also proposed.

    However, the correlation peak of the timing metric exhibits a plateau which causes

    large variance for the timing estimator [21]. Based on a training preamble having

    the same structure as the one used in the Schmidl-Cox’s estimator [19], reference [20]

    showed that the performance of the MMSE (Minimum Mean-Square Error) and ML

  • 44

    (Maximum Likelihood) timing estimators still perform unsatisfactorily.

    Similar to the signaling set-up adopted in [19], Coulson used two repeated m-

    sequences as a training symbol [24]- [25]. However, the proposed time synchronization

    algorithm is likely to fail in the presence of large carrier frequency offsets, and presents

    high implementation complexity due to the matched filtering. We remark that a

    training symbol with the same structure as the one proposed in [24] is exploited

    in [26] to develop reliable frequency and time acquisition schemes. However, the

    proposed time synchronization algorithm is also sensitive to large frequency offsets.

    By adopting a structured training symbol of the form [+B +B -B -B], Bhargava

    et.al. proposed a coarse timing delay estimator that outperformed Schmidl-Cox’s es-

    timator [21]. However, reference [21] does not provide any detailed insight or analysis

    pertaining to the features of this estimator. Inspired by the signaling set-up proposed

    in [21], this chapter aims to develop reliable and reduced complexity coarse frame

    and carrier frequency acquisition schemes that exploit a structured training symbol

    of the form:

    [±B ± B ± B ± B] , (4.1)

    where B stands for a sequence of N/4 training samples with constant variance (power)

    (e.g., an m-sequence), and it can be generated with good approximation by using an

    N/4-point IFFT of an m-sequence. Principally, any signal with constant envelope in

    the time domain and a bandwidth similar to the OFDM data symbol can be used as

    a training symbol. It is found that among all the signaling set-ups (4.1), the training

    symbol:

    [+B + B − B + B] , (4.2)

    leads to timing acquisition schemes that exhibit the best detection properties in terms

    of lower false detection probability and higher correct acquisition probability. In

  • 45

    addition, it turns out that the negative term -B present in (4.2) can be placed in

    any location with no loss in performance1. By exploiting the structured training

    symbol (4.2), this chapter proposes robust acquisition schemes for time delay and

    carrier frequency offset for OFDM systems that operate in additive white Gaussian

    noise (AWGN) and frequency-selective channels. The proposed time synchronizer

    offers more accurate estimates than the estimators [19] and [21]. It is also found that

    the performance of the proposed time and frequency offset estimators is nearly the

    same as [24]. In addition, the proposed estimator requires a reduced implementation

    complexity and is more robust to large frequency offsets with respect to (w.r.t) [24].

    The rest of chapter is organized as follows. In Section B, we describe the signal

    model and introduce some modeling assumptions. In Section C, an optimum maxi-

    mum likelihood (ML) estimator is derived for the continuous transmission scenario.

    We modify this ML estimator to a sub-optimum estimator with reduced complex-

    ity, which is shown to exhibit robust performance to fading channel for both burst

    and continuous transmission scenarios. A theoretical performance analysis study is

    conducted in Section D. Finally, Section E describes computer simulations that illus-

    trate the advantages of the proposed estimator and that corroborate the theoretical

    performance analysis performed in Section D.

    B. Signal Model

    The OFDM baseband signal is generated by the IFFT-transform:

    x(n) =1√N

    N−1∑

    k=0

    skej2πkn/N , −L ≤ n ≤ N − 1 , (4.3)

    1All of them are similar to the bark code with length 4

  • 46

    where sk represents the data sequence modulated on the kth subcarrier, which may

    assume any modulation format (such as QAM or PSK), and is independently and

    identically distributed (i.i.d.) with zero mean and variance E{|sk|2} = σ2s . Let N ,

    L, and 1/Tu = 1/(NTs) denote the number of subcarriers, the length of cyclic prefix

    (guard time), and the subcarrier spacing, respectively. Normally, the length of cyclic

    prefix L is selected to be not more than N/5, which can be interpreted as a 1-dB

    signal-to-noise ratio (SNR) loss introduced by the cyclic prefix [17, p. 46]. Herein,

    without loss of generality (w.l.o.g.) the value L = N/8 is adopted.

    At the beginning of our study, we assume a flat fading channel model to derive the

    maximum likelihood estimator. Later it will be shown by computer simulations that

    the proposed estimator works well for frequency-selective channels, too. Assuming

    the sampling period Ts = Tu/N , the received signal samples can be expressed as:

    r(n) = α(n)x(n − ϑ)ej(2πfnn/N+θ) + w(n) , (4.4)

    where ϑ stands for the timing offset, α(n) is the channel amplitude, fn := feTu =

    feNTs is the normalized carrier frequency offset, θ denotes the phase offset, and w(n)

    denotes the samples of a zero-mean complex white Gaussian noise random process

    with variance E{|w(n)|2} = σ2w and is assumed independent w.r.t x(n). In slow-

    varying channels, we can assume α(n) to be a constant α over the duration of several

    OFDM symbols. The signal-to-noise ratio is represented in terms of the variable

    SNR:=α2σ2s/σ2w.

    After coarse frame and carrier synchronization is achieved, the receiver discards

    the cyclic prefix, and the modulated symbol stream {sk} can be recovered by means

    of an FFT-operation. Due to the presence of cyclic prefix, small (fractional) time

    offsets ϑ in magnitude less than the interval (L) spanned by the cyclic prefix cause no

    ISI or interchannel interference (ICI) [49]- [50]. The time offset induces a phase offset

  • 47

    exp (−j2πϑn/N) on the nth subcarrier, which can be corrected using channel esti-

    mation techniques. Therefore it is enough to estimate the start of training sequence

    within one sample period.

    This chapter focuses on the estimation of fn (packet detection) and ϑ. Although,

    α and θ are unknown to the receiver, their estimation can be avoided via differential

    encoding/decoding or can be obtained wi